Skip to main content

Full text of "Equilibrium adsorption on a random site surface"

See other formats


Equilibrium adsorption on a random site surface 



J. Talbot 1 , G. Tarjus 2 and P. Viot 2 
1 Department of Chemistry and Biochemistry, Duquesne University, Pittsburgh, PA 15282-1530 
2 Laboratoire de Physique Theorique de la Matiere Condensee, 
Universite Pierre et Mane Curie, 4, place Jussieu, 15252 Paris Cedex, 05 France 

We examine the reversible adsorption of spherical solutes on a random site surface in which the 
adsorption sites are uniformly and randomly distributed on a substrate. Each site can be occupied 
by one solute provided that the nearest occupied site is at least one diameter away. The model is 
characterized by the site density and the bulk phase activity of the adsorbate. We develop a general 
statistical mechanical description of the model and we obtain exact expressions for the adsorption 
isotherms in limiting cases of large and small activity and site density, particularly for the one 
dimensional version of the model. We also propose approximate isotherms that interpolate between 
the exact results. These theories are in good agreement with numerical simulations of the model in 
two dimensions. 

PACS numbers: 



I. INTRODUCTION 

Heterogeneity often plays an important role in various 
adsorption processes and its presence may profoundly 
modify the adsorption isotherms and other thermody- 
namic properties compared to the homogeneous situa- 
tion. Consequently, numerous articles [J 0], reviews [1,0] 
and monographs 0, @ have been published that describe 
experimental, theoretical and numerical studies in the 
area. The disorder may originate from the energetic or 
structural heterogeneity of the substrate or from the ad- 
sorbate species. Heterogeneity may also result if the ad- 
sorbed molecules are large enough so that multisite ad- 
sorption becomes possible 0j- 

In a recent article [|[ , we investigated the properties of 
a lattice model of adsorption on a disordered substrate 
that can be solved exactly (See also Refs.[9|, G3). We 
showed that there exists an exact mapping to the system 
without disorder in the limits of small and infinite activ- 
ities and we exploited this result to obtain an approxi- 
mate, but accurate description of the disordered system. 

For continuous systems, structural disorder may be 
represented by the random site model (RSM) in which 
adsorption sites are uniformly and randomly distributed 
on a plane [ill 0- The molecules, represented by hard 
spheres, can bind to these immobile sites. Steric exclu- 
sion is expressed by the fact that a site is available for 
adsorption only if the nearest occupied site is at least one 
particle diameter away. In addition, adsorption energy is 
assumed equal for each adsorbed molecule. Therefore, 
the disorder of this model is characterized by the dimen- 
sionless site density, p Sl only. The degree of complex- 
ity increases drastically because steric effects, which usu- 
ally dominate the adsorption on continuous surfaces, are 
modified by the local disordered structure of adsorption 
sites. This model may be appropriate for the reversible 
adsorption of proteins on disordered substrates (ljj, [Tij . 

Oleyar and Talbot [12] examined the reversible adsorp- 
tion of hard spheres on the 2D RSM and proposed an 
approximate theory for the adsorption isotherms based 



on a cluster expansion of the grand canonical partition 
function. Although successful at low site densities, the 
quality of the theory deteriorates rapidly with increasing 
site density and fails completely above a certain density. 

In sections II and III we develop a general statistical 
mechanical description of the RSM model and we confirm 
the intuitive result that in the limit of large site density 
the system maps to hard spheres adsorbing on a continu- 
ous surface. Then, for finite site density, we prove that for 
the one-dimensional system in the limit of infinite activ- 
ity, there is a mapping to a hard rod system at a pressure 
[3P = p s . We also present an argument supporting the 
validity of this result in higher dimensions. Note that 
when the activity is "strictly" infinite, desorption is no 
longer possible and the adsorption process is irreversible. 
In this case the model has an exact mapping to the Ran- 
dom Sequential Ad sorp tion (RSA) of hard particles on a 
continuous surface [ilj . 

In sections IV and V we propose approximate theoret- 
ical schemes for the adsorption isotherms in one and two 
dimensions that interpolate between the limits of small 
and large activities for a given site density p s . Compari- 
son with simulation results shows that these approaches 
are a considerable improvement over the cluster expan- 
sion. 

For completeness we note that in addition to its appli- 
cation to adsorption, the model is also interesting because 
of its relationship to the vertex cover problem (la. [iH] . A 
vertex cover of an undirected graph is a subset of the ver- 
tices of the graph which contains at least one of the two 
endpoints of each edge. In the vertex cover problem one 
seeks the minimal vertex cover or the vertex cover of min- 
imum size of the graph. This is an NP-complete problem 
meaning that it is unlikely that there is an efficient algo- 
rithm to solve it. The connection to the adsorption model 
is made by associating a vertex with each adsorption site. 
An edge is present between any two vertices (or sites) 
if they are closer than the adsorbing particle diameter. 
The minimal vertex cover corresponds to densest particle 
packings. Weight and Hartmann[l5|, [l6| obtained an an- 



2 



alytical solution for the densest packing of hard spheres 
on random graphs, but the existence of the geometry in 
adsorption processes implies that the machinery devel- 
oped to describe adsorption on random graphs cannot 
be used in RSM models. 



II. STATISTICAL MECHANICS OF THE 
RANDOM SITE MODEL 

The adsorption surface is generated by placing n s 
points, representing adsorption sites, randomly and uni- 
formly on a substrate, either a line in ID or a plane in 
2D (the boundary conditions are irrelevant in the large n s 
limit). Spheres of diameter a may bind, centered, on an 
available adsorption site. A site is available if the nearest 
occupied site is at least a distance a away. Two points 
are connected, and therefore cannot be simultaneously 
occupied, if they are closer than a. 

The positions of the n s sites are denoted by Ri where 
i is a index running from 1 to n s . The sites are quenched 
during the adsorption-desorption process. The adsorbed 
phase in equilibrium with a bulk phase containing ad- 
sorbate at activity A can be formally described with the 
grand canonical partition function: 



s(a,{rj)=i+^- / ... /dr™n( i +/«')n^) 

n=\ ' ^ ^ i>j i=l 

where the microscopic density of sites r\ (r) is given by 



» 7 (r)=£«(r-R i ), 



(2) 



A = exp(/3/z) is the activity, denotes the position of 
sphere i and /y is the Mayer f-function which is equal 
to —1 if spheres i and j are closer than a and other- 
wise. Eq. (fl]) applies to a particular realization of the 
quenched variables {Ri}. In order to average over disor- 
der, it is necessary to take the average not of the partition 
function, but of the logarithm of the partition function 
By using standard rules of diagram theory [l8j|, one 
obtains that 

RSJ-E— / dv n U n {v 1 ,r 2 ,...,v n )J[^i) (3) 
n=l n " J i=l 

where the bar means that the average is taken over dis- 
order, U n (ri, r 2 , r„) denotes the Ursell function as- 
sociated with the Mayer f-functions of hard particles, 

II, ,ii • fvM- 

Let us denote the probability of finding adsorption sites 
at positions Ri, i = 1, n s , as P(Ri, R2, ...R s ). We will 
assume that the positions are uncorrelated so that 



P(Ri,Ra,...R.) = fJP(Ri) 



(4) 



and will consider a Poissonian distribution of points, 
P(R) = 1/A. The average of the site density ry(r) is 
given by 



V(r) =/-/ n dR * P ( R i' R2 '- R sM r ) 

J •* i=i 

= Ps (5) 



where p s = n s /A is the site density of the particles in a 
system of area A (length in one dimension) . 

We show in Appendix A that the average of the loga- 
rithm of the partition function over disorder can be writ- 
ten as 



ln(S)=ln(H*(« = Ap a )) + 0(l/p,) 



(6) 



where 



ln(S*(z)) = E^/ dr"t/„(n,r 2 ,...,r n ) (7) 



i.e., ln(S*(z)), is the partition function of hard spheres on 
a continuous surface at an activity z = Xp s . This result 
shows that when the site density p s goes to infinity and 
that the activity A goes to 0, with the constraint that 
the product Xp s remains finite, the Random Site Model 
maps to a system of hard particles in continuous space. 

The number density of adsorbed molecules can be com- 
puted directly from the partition function: 



p{X > Ps) = A-(^ 



(8) 



with again z = Xp s . 

By using Eq. (|A8[) given in the Appendix, one obtains 
to the second-order in l/p s 



p(\,p s ) = p*{z) 



P * {Z) dP * [z) '-0(1/ ^ (9) 



Ps+p*(z) dz 



where p*(z) is the number density of the hard sphere 
model in continuous space. Since p*{z) is an increas- 
ing function of the activity z, it is an upper bound for 

The expansion of lnS and p(X,p s ) in powers of the 
activity when A — > can also be generated from the 
above equations in straightforward way. 



III. THE LIMIT OF LARGE ACTIVITY 

A. A mapping with the (homogeneous) equilibrium 
hard sphere model 

The limit of large activity, A — > 00, is expected to lead 
to the maximum density of adsorbed spheres for a given 
density p s of adsorption sites. Indeed, this limit com- 
bines the presence of a relaxation mechanism, which is 



3 



induced by the infinitesimally small but non-zero desorp- 
tion process that allows a sampling of hard-sphere con- 
figurations, with the guarantee that no sites left open for 
adsorption will stay empty. The one-dimensional model 
is then amenable to an exact solution in the limit A — >• oo. 
Interestingly, it maps onto an equilibrium system of hard 
rods in continuum (ID) space at the same density (which 
is of course a function of p s ) 

The adsorbed density p(p s , A — > oo) = p ma , x (p s ) in the 
ID case can be obtained exactly with the following sim- 
ple probabilistic argument. Assume that a given site is 
occupied. Then all sites that lie within a distance a from 
this site must be unoccupied. In the the optimally packed 
system the next site beyond this must be occupied. The 
average distance from an arbitrary point to the first site 
is 1/ps giving the average distance between two occupied 
sites as a + 1/ p s . The average coverage in the maximally 
occupied system is thus 

Pmax(ps) = — — • (10) 

1 + Ps<J 

In the following we set a = 1. When the site density is 
low, p s << 1, one recovers the independent site approxi- 
mation (Langmuir model), p max = Ps- As the site density 
increases Eq. (TTU|) shows that there is a continuous in- 
crease of /O max and that a closed packed configuration is 
obtained as p s approaches infinity 

The above argument can be generalized to describe 
the correlation functions, or more conveniently in this 
ID system, the gap distribution functions. Specifically 
let F{x) denote the probability density associated with 
finding a gap of size between x and x+dx. In an optimally 
packed configuration, the probability to find a gap of 
length x is related to the probability to find the first site 
at a given distance x (say to the right) of an arbitrary 
point. For a Poissonian distribution of sites this simply 
leads to 

F max (x-p s )=p s e- p ° x . (11) 

The reasoning is easily extended to multi-gap distribu- 
tion functions, F[x\, x%), F(x±, X2, £3), where two suc- 
cessive gaps are neighbors in the sense that they are sep- 
arated by a single particle. One then shows that all 
these higher-order gap functions factorize in products 
of one-gap distribution functions, e.g. F max (xi, x%) — 
-Pmax(a^i)-Fmax(^2). The outcome of this probabilistic ar- 
gument is that the ID random site model is equivalent 
to an equilibrium system of hard rods on a (continuous) 
line at the pressure 

0P = Ps- (12) 

Indeed, from the known equation of state of the hard rod 
fluid Hfll one has (3P = p/(l - p), i.e. p = @P/(1 + (3P), 
which corresponds to Eq. (flU)) after insertion of Eg. (fT2l -. 
and F(x) = (3Pe~P Px , which reduces again to Eq. (fTTjl . 
The multi-gap distribution functions are also given by 
products of 1-gap distribution functions, which completes 
the proof of equivalence. 



Extension of the above arguments to higher dimensions 
is far from straightforward. First, in d — 2 and higher, 
one may encounter at high adsorbed density phase tran- 
sitions to ordered, crystalline-like, phases. Second, the 
simplicity of the reasoning in terms of gaps character- 
ized only by their length is lost when one leaves the one- 
dimensional case. For these reasons, we have not been 
able to develop a rigorous demonstration of a mapping 
between the 2D RSM and the (homogeneous) hard-disk 
system at equilibrium at the same density when the ac- 
tivity A goes to infinity. 

To nonetheless make some progress, let us consider the 
nearest neighbor radial distribution function H(r) intro- 
duced by Torquato |2l]. H(r) (and also used in the con- 
text of irreversible adsorption models [22|,[23[ is the prob- 
ability density associated with finding a nearest neighbor 
particle center at some radial distance r from the refer- 
ence particle center. This somewhat generalizes the 1- 
gap distribution function to any spatial dimension. For 
a Poissonian distribution of particle centers of density p s 
in d — 2, one finds that 

H(r) = 2irrp s e- 7Tr2p ° (13) 

No exact formula exists for H(r) in a hard disk fluid. Us- 
ing the results of reference [21| one can, however, demon- 
strate that in the large r limit at an equilibrium pressure 
P it is given by 

H{r) ~ 2Trrf3Pe-^ Tr2 > 3P (14) 

The reasoning now goes as follows. Consider a maxi- 
mally occupied configuration (associated with the limit 
A — > 00) in the 2D RSM with an adsorption site density 
p s and consider a given adsorbed particle. All sites within 
a radial distance a of the central occupied site must be 
unoccupied. H(r) is associated with situations such that 
the nearest adsorbed particle center is at a radial dis- 
tance r of the reference one. Therefore, there should be 
no adsorption sites in the region delimited by the two 
circles of radius a (inside) and r (outside). Actually, this 
statement is not quite right: in 2 dimensions, there may 
be an exclusion effect due to other particle centers at a 
distance > r of the central site. To be more rigorous, the 
outside circle delimiting the region with no adsorption 
sites should be (at least) of radius r — a. When r — ► 00, 
one can neglect the contribution of width a due either to 
the disk of radius a centered on the reference site or to 
the shell of width a located between r — a and r. When 
r — > 00, the probability that, given that a reference par- 
ticle is centered at the origin, a spherical region of radius 
r is empty of adsorption sites is given by exp(— p s 7rr 2 ), 
so that asymptotically H max (r) goes as in Eq. (p~3|) . i.e. 

ffm«(r) ~ 2-Krpse-^P' (15) 

Comparison with Eq. (TT4"|) tells us that it is the same 
asymptotic behavior as that of an equilibrium hard disk 
system in the plane with /3P = p s . This is the same 



4 




1/ln(X) 



FIG. 1: Simulated adsorption Isotherms. p s = 1,2,4 top to 
bottom. The dashed lines show the predictions of Eq. (|10|l . 

result as found in d = 1 (see above), except that the 
density p and pressure P are no longer related by a simple 
analytical expression. 



B. Numerical verification of the suggested mapping 

We confirmed Eq. (JTUJ) in two ways. The first generates 
configurations of maximum density directly. A number 
n s points are distributed uniformly and randomly in the 
unit interval. The first site is occupied by a rod of length 
a = l/n s . Each site to the right is checked in order until 
one is found that is at least a distance a from the occu- 
pied one. This site is occupied and the process iterated 
until all sites have been accounted for. A number of av- 
erages over different configurations of sites is performed. 
We also verified Eq. (fT0|) by determining the adsorption 
isotherms for different values of the site density p s and 
taking the limit A — > oo: See Fig. [TJ 

For the hard rod fluid at equilibrium the gap distribu- 
tion function is given exactly by: 

F(p,x) = -^ex V (--^). (16) 
1-p 1 - p 

The gap distribution function calculated for the dens- 
est configurations of the random site model is in agree- 
ment with the predictions of Eq. (|16p using a density 
computed from Eq. (JTUJ), confirming that the configu- 
ration of rods on the random site surface has the same 
structure as the equilibrium hard rod fluid at the same 
density (see FigJ^]). This is consistent with the behavior 
of a related lattice model for which we showed that there 



FIG. 2: Gap distribution function F ma x(a:) for configurations 
of hard rods at infinite activity on the random site surface 
with p s = 0.7,1,2.0,4.0 from right to left, bottom. The 
dashed lines are the predictions of Eq.fTTJ The solid lines 
correspond to the simulation results. 



is an exact mapping to the system without disorder at 
infinite activity [8j. 

In order to use Eq. (|T5|) to estimate the maximum cov- 
erage of the RSM in two-dimensions, we combine it with 
the approximate equation of state for hard disks on a 
continuous surface, which was proposed by Wang [241]: 



— = 1+ ; (D+2a9+4b9 2 +8c9 3 +16d9 i +32e9 5 ) 

p 1-0/00 

(17) 

where 9 = Tia 2 p/A is the coverage, 9$ — 0.907.. is 
the coverage of a hexagonal close-packed configuration, 
D = 4.08768, a = 1.25366, b = 0.46051, c = 0.152797, d = 
0.04412, e = 0.00929.. For a given value of the site den- 
sity, p s , Eq. (|17p is solved numerically for 9. It is conve- 
nient to introduce the dimensionless site density 

a = ^a 2 p s (18) 

The results, shown in Fig. [3] are in excellent agreement 
with the simulation results for the entire range of a. This 
should be compared with cluster expansion to second or- 
der that gives good predictions only for a < 0.3[12]. Note 
that the approximate equation of state, Eq. (|17p . does not 
include the possible presence of a phase transition. 



5 



be expressed as 




FIG. 3: Maximum coverage of the RSM as a function of the 
dimensionless site density. The symbols show the simulation 
results, the dashed line is the cluster approximation to second 
order and the solid line shows the predictions of Eqs. (|12|) 
and (fT7)) . 



InS = Ni In Hi + N 2 In S 2 + N 3a In E a + N 3b In E 3b + 

(19) 

where N — XiN s is the number of clusters of type i 
and i = (n, a) with n being the number of sites in the 
cluster and a characterizing when necessary the subclass 
of clusters (see FigQJ. This simple expansion is possible 
because adsorption on a given cluster does not affect any 
of the others. The adsorption isotherm is then 



pA = XN X 



which gives 



In Si 



XN 2 



d In Ej 
~0X~ 



(20) 



P = \p s (xi 

3 + 2A 
+ 1 + 3A + A 2 



1 + A 



X 3a 



1 + 2A 
3 

1 + 3A 



3'2 
X 3 b 



(21) 



In one-dimension, exact expressions can be obtained 
for the first few clusters following the approach of Quin- 
tanilla and Torquato [25[: 



1+X 



1+3X+X 1 



A 



1+2X 



1+3X 



FIG. 4: Lowest order clusters of adsorption sites. A solid line 
connects two sites that are closer than a and a dashed line 
indicates that two sites are further apart than a. Triplets of 
type "3a" (left) and "3b" (right) are shown left to right in 
the second row. The expression to the right of the clusters is 
the corresponding grand canonical partition function for the 
adsorbed particles. 



IV. RSM IN ONE DIMENSION 
A. Low site density expansion 

To provide a description of the adsorption isotherms 
at finite activity, it is useful to consider an expansion 
in the density of adsorption sites. Assuming that the 
adsorption surface consists of isolated clusters of sites, 
the (averaged) logarithm of the partition function may 



x x = exp(-2p s ) (22) 

x 2 = exp(-2p s )(l - exp(-p s )) (23) 

x 3a = cxp(-3 / 9 s )(exp(- / 9 ;5 ) + p s - 1) (24) 

x 3b = exp(-2j(? s )(l - (1 + p a ) exp(-p s )) (25) 

where x 2 , for example, is the number of clusters with 
two linked sites (per adsorption site). At the triplet level 
we need to distinguish between two subclasses of clusters 
shown in Fig. |H In type "3a" two of the sites can be 
simultaneously occupied, while in type "3b" only one of 
the sites can be occupied since all are mutually closer 
than the particle diameter. 

The predictions of Eq. (|2"Tj) are compared with the sim- 
ulation results in Fig. [5l The expansion to third order 
provides a good description of the isotherm for p s = 0.2, 
but the quality deteriorates rapidly for larger site den- 
sity. Eq. (|2ip is unable to predict the correct limit when 
A — > oo. This disagreement is more pronounced when p s 
increases: for instance, when p s = 1, the expansion fails 
completely, because the predicted density is lower than 
for p s = 0.5. 

Since the above expansion is limited to small values of 
p s , we simplify Eq. (|21[) by performing a series expansion 
in p s , giving 



2A 



+ A (1 + A)(1 + 2A)' 
A 2 (12A 2 + 29A + 9) 



2(1 + A)(l + 2A)(1 + 3A)(1 + 3A + A 2 ) ' 



0(p 3 s )) 



(26) 



6 



We note that Eq. (|26|) can be expressed as 



p(Ps, A) = 



1=1 



1 + A 



+ F l (p s ,X)\ p[ (27) 



where Fi(p s , A) represents the remaining terms in the ex- 
act expansion and, consistent with these terms, has the 
property that Fi(p s ,X) — > when A — > and A — > cxx 
This property has been verified order by order for a ran- 
dom lattice model[§j, and only for the three first orders 
for this model. If we partially resum the series, neglecting 
the remaining terms, we obtain that 



Xp s 



1 + A(l + Ps ) 



(28) 



Although the site density expansion, Eq. (I26|) . is valid 
only for small p s , the partial resummation leads to an 
expression for p that is exact in the limit of very large 
activities A, whatever the value of p s . It thus provides a 
sensible approximation based on the site density expan- 



Isotherms given by Eq. (f2"8")) are plotted in Fig. [5] (dot- 
ted curves). Although for very low site density the third 
order expansion gives a better estimation, the quality of 
Eq. P5|) is significantly better for p s — 0.5, and it always 
gives the saturation density exactly. For larger values of 
the site density, p s > 1, the approximation overestimates 
the adsorbed amount for intermediate activities. This 
is expected since the neglected terms are non-zero for 
intermediate bulk activity. Attempts to obtain a bet- 
ter resummation of the series expansion, Eq. (I26|) . were 
fruitless. Therefore, we follow the approach that we used 
in our study of the model of dimer adsorption Q, by in- 
troducing an effective activity. 



B. Effective activity approach 

Despite the exact mapping found in the infinite ac- 
tivity limit, no simple reasoning can be used to obtain 
exact results at finite bulk activity. We therefore investi- 
gated several approximate methods, the most successful 
of which is based on an effective activity. 

The equation of state of the hard rod fluid on a con- 
tinuous line is [20( 



PP = 



1 — per 



(29) 



According to the results in Section III, the hard rod fluid 
at a pressure f3P = p s has the same density (obtained by 
inverting Eq. (|2"9"]) ) as the densest configuration of hard 
spheres on the random site surface, Eq. (fTU)) . We seek a 
generalization of this mapping for an arbitrary value of 
the bulk phase activity, i.e. an effective activity A e fi such 
that the density of adsorbed rods in the RSM, p(p s , A) is 
given by 




ln( X) 



FIG. 5: Adsorption isotherms versus activity A for several site 
densities: p 3 = 0.5, 0.3, 0.2 from top to bottom. The dash- 
dotted and dotted lines show the predictions of Eq. [21] (exact 
expansion to third order in p s ) and Eq. 1281 respectively. The 
solid lines show the simulation results. 



where p*(X) is the density of rods on a continuous sub- 
strate at an activity A. This is given exactly by 



P*(A) 



MA) 

i+L w {\y 



(31) 



where L w (x) , the Lambert-W function, is the solution of 
x = L w (x) exp(L w (x)). The inverse relation is 



A = 



P 



; exp( — 



(32) 



From the exact result, Eq. (fTTJ|) , we have that A e ff(A = 
oo) = p s e Ps . Taking Eqs. (|31[) as a mere definition of 
A e ff(A,p s ), we have computed it from the simulated ad- 
sorption isotherms. Results are shown in Figs. [5][5] and 
will be used as a reference for testing the validity of var- 
ious approximations. 

Following the method introduced in the study of the 
lattice modelfl], we attempt to construct simple expres- 
sions that can reproduce the numerical results, when 
available and describe the isotherms for a wide range of 
parameters p s and A. The effective activity as a function 
of the activity A and of the density site p s can be written 
as 



1 = f(Ps,X) | 1 
A e ff p s A p s eP° 



(33) 



p(p s ,X) = p*(X cS (X,p s )), 



(30) 



where f{p s ,X) is a function to be determined. We 
know that at small A, the exact behavior is given by 



7 




200 



150 



\a(X) 




100 



10 12 



FIG. 6: Effective activity A e a as a function of the logarithm 
of the activity In A with p s = 1. The solid line shows the 
simulation result, calculated from the density by using Eq. 
(1321) . The predictions of Eq. J33} combined with Eq. (J34J) and 
Eq. (l36|l are shown as dotted and dashed lines, respectively. 
They are almost indistinguishable. The horizontal dashed 
line shows the exact limiting value of p s e Ps ■ 




10 12 



FIG. 7: Same as Fig. [6] except that p s = 2. The prediction 
of Eq. (|36p is now more accurate than that of Eq. (|34|l . 



A c ff = p s A, which imposes that f(p s , A) goes to 1. Con- 
versely, when A is large, the effective activity A e ff be- 
haves as p s e Ps + O(A), which leads to the constraint that 
liniA^oo f(Psi A) = A(p s ) where A(p s ) is an unknown 
function of p s . If A(p s ) remains close to one, the sim- 
plest choice consists of choosing 

/(p s ,A) = l (34) 

for all values of A. While this simple choice gives a 



FIG. 8: Same as Fig. except that p s = 4. Eq. (J34} fails to 
reproduce the simulation data for intermediate value of ln(A) 
whereas the prediction of Eq. (|36p remains accurate. 



fair agreement with the simulated values when p s = 1 
(see Fig. the situation deteriorates for larger values 
of p s (see Figs. [?)■ [8]). However, contrary to the one- 
dimensional lattice model, we do not know the exact 
asymptotic behavior of f{p s ,\) for large A. In order 
to improve the description, we consider a homographic 
function of A verifying the exact behavior, when A — * 
and A — > oo, 



f(Ps,X) 



A(p s )X + B( Ps ) 
A + B(p s ), 



(35) 



where A{p s ) and B{p s ) are unknown functions. Exami- 
nation of the series expansion in the limit of infinite bulk 
activity suggests that B(p s ) should grow as e Ps in order 
to give the right behavior. In the absence of additional 
criteria, we have chosen A(p s ) — p s and B(p s ) = p s e p " , 
which gives 



/(P., A) 



A) 



A + p s 



(36) 



Note that when p s = 1, this function gives f(p s ,X) = 1. 
Figs. [BJ- H] show that the quality of the approximation 
given by Eq. (f3"3")) and Eq. (fM)) deteriorates with in- 
creasing p s . This is a consequence of the fact that the 
asymptotic approach to the saturated state is not cor- 
rectly captured with this simple approximation. A better 
agreement with simulations is obtained by using Eq. (|33|) 
and Eq.® for p s = 2 and p s = 4 (see Figs. [7] and HJ. 

The purpose of this exercise is ultimately to obtain an 
approximate, but accurate, description of the adsorption 
isotherms. This is done by substituting Eq. (|33p in Eq. 
(l3"Tj) . Fig. shows a comparison of these estimates with 
the simulation results. The effective activity approach is 



8 




10 12 




0.01 



FIG. 9: Adsorption isotherms versus activity A for several 
site densities: p s = 1,2,4, predicted by the effective activity 
approach, Eqs. (|32|l and (|31[1 by using Eq. (|34p (dotted lines) 
and combined with Eq. (|36[) (dashed lines). The solid lines 
show the simulation results. 



a considerable improvement over the cluster expansion, 
whose accuracy is restricted to small density site p s < 0.5 
(c.f. Fig. [5]) and allows one to describe adsorption even 
when p s > 1. 



C. Structure 

We have also examined the structure of the hard-rod 
configurations at finite activity. As in the case of the lat- 
tice modelQ, we do not expect an exact mapping to the 
homogeneous system at the same density. Nevertheless, 
as shown in Fig. 1 1 01 the distributions computed from 
the simulation are very well described by Eq. (|16[) with 
a density equal to the equilibrium isotherm value. 



V. TWO-DIMENSIONAL MODEL 

We construct approximate isotherms in the same way 
as for one-dimensional model, i.e. by introducing an ef- 
fective activity as a function of bulk activity. 

We first consider the homogeneous hard-disk system. 
We determine the Helmholtz free energy per particle by 
integration along the isotherms: 



(37) 



The excess chemical potential can then be obtained from 



FIG. 10: Gap distribution function for configurations of hard 
rods at finite activity on the random site surface with p 3 — 10. 
The dashed lines are the predictions of Eq. [16] with a density 
given by the equilibrium isotherm value and the solid lines 
show the simulation results. A = 1,5,50,1000 from left to 
right in the bottom part. 



By inserting the approximate equation of state, Eq. (JT7J) 
in Eq. (|3"7) . and substituting the result in Eq. ([581) . one 
obtains for the activity, A = exp(/3/i) the relation 

OR 

ln(A) = ln(40/7r) e(9 5 - 20 d<9 4 c6 3 - 6 b9 2 - 4 at 

5 3 



-Din 1 - 



2 V3\ 



DV36* 
7T - 2 V30 ' 
(39) 



where, we recall, 9 = ira 2 p/4. 

The isotherm for the 2D RSM can be calculated by 
following a procedure similar to that used in the one- 
dimensional case. By means of Eq. (fT2"|) , one obtains the 
saturation coverage when the bulk activity is infinite, #oo . 
Then, by inserting this quantity in Eq. (|39|) , the effective 
activity A^ = A e ff(#oo) is obtained. 

The counterpart of Eq. (|33[) for the two-dimensional 
system is now 



1 

Acff 



1 



1 

A off 



(40) 



j3pL ex = (3f ex + i 



de 



(38) 



where, in the absence of more information, we have set 
/(P.,A) = 1. 

Therefore, at a given bulk activity A and site density 
p s , one obtains an effective activity by using Eq. (|4"0")l . 
and by inserting the effective activity in Eq. (IST))) . one 
deduces the corresponding coverage in the 2D RSM. 



9 




FIG. 11: Adsorption Isotherms for dimensionless site densities 
a = 1,2,4 from bottom to top. The dashed lines show the 
predictions of Eg. (|39p combined with Eg. (|40p . The solid lines 
show the simulation results obtained with 20 realizations of 
n s = 2000 sites. 



Fig. [TT] compares the simulation results and the ap- 
proximate isotherms for different values of the site den- 
sity. While the agreement is not perfect, the scheme 
represents a major improvement compared to the pertu- 
bative approach (third-order density expansion) for de- 
scribing situations where the density site is moderate to 
high and allows one to develop new approximate isotherm 
equations. 



VI. CONCLUSION 

We have presented a theoretical and numerical study 
of the reversible adsorption of hard spheres on randomly 
distributed adsorption sites. Unlike the case of irre- 
versible adsorption, there is no simple mapping between 
the RSM and a system of hard spheres on a homoge- 
neous surface. We have, nonetheless, been able to ob- 
tain some exact results in various limits: small and large 
bulk activity and small and large site density. We have 
proposed an effective activity approach, which interpo- 
lates between the known behavior, to obtain an approx- 
imate description for intermediate situations. The ad- 
sorption isotherms predicted by this approach are in ex- 
cellent agreement with simulation results. In the two- 
dimensional case, we have not investigated phase transi- 
tions in detail, although there does appear to be a solid- 
like phase for sufficiently high site density and activity. 
It will also be interesting to investigate reversible ad- 
sorption on a substrate when a second source of disorder 



is present: distribution of adsorption site energies. The 
present approach should apply to this case as well. 



APPENDIX A: RANDOM SITE MODEL: 
BEYOND THE CONTINUUM LIMIT 

The sites being uncorrelated, Eq.Q, one eas- 
ily obtains the connected n-site correlated functions 

^(ri)7/(r 2 )...r?(r„)^ which read 

(p{riHr2)-v(r n )) c = Ps Jpfa - ( A1 ) 

The n-site correlation functions which appear in Eq. |T]) 
can be obtained from the above connected functions by 
using an expansion in the (mean) site density: 

77(n)77(r 2 )...77(r n ) = p'l + pT 2 X (WM^f) c 

i<j 

i<j<k 
i<j<k<l 



i<3 



i<k<l 



(A2) 



The expression of the higher-order terms rapidly becomes 
tedious. 

On the other hand, the n-body connected density func- 
tions of a homogeneous hard sphere system at activity z 
are given by functional derivatives of the the logarithm 
of grand-canonical function (see Eq.(j3|) 



pW*(r ir . 2 ...r p |* 



SP ln(S*(z(r))) 



z{r)—z 



X 



(n — p)\ 



n—p 



8z(yi)8z(y 2 )...8z(yp) 
J dr p+1 ...dr n U n (r 1 ,r 2 ...r p ,r p+1 ...r n ). 

(A3) 

By inserting Eq. (jA2| in Eq.© and by using Eq.fA3}, the 
disorder averaged logarithm of the partition function of 
the RSM has the following expansion: 

ni^ = ln(H*(^) + ^-|dr 1 p( 2 )*(r 1 ,r 1 |^ 

+ 3!7 y*ipi 3) *(n,ri,nW 

+ ^3 J *i^ 4) *(ri,ri,ri,ri|z) 

+ 4^2 J dridr 2 ,9* (4) (ri,ri,r2,r2|z) 

+ ... (A4) 



10 



For hard spheres, the non-overlapping property leads 
to a simple expression for the n-body connected density 
function 



(rx.n.^n.nl^) = {-if^ip- i)\p*(z) p . (A5) 



A straightforward but more tedious calculation allows 
one to obtain 

p( 4 )*(r 1 ,r 1 ,r 2 ,r 2 |z) = V(z) 2 (0 ( 2 )*(r 1 ,r 2 )-2p( 2 )*(r 1 ,r 2 ) 2 

(A6) 

By inserting Eqs. (IA5j) and (IA6j) in Eq. (IA4j) . one obtains 
that 



MS) _ ln(3*(z)) , ( p*(zY , p*{zY p*(z) 

A 
p*{zf 

2p: 



A ' \ 2 Ps 3 P 2 Apl 
2 dr(2h*(r\z) - (h*(r\z))) 2 (A7) 



where h*(r\z) is the radial pair correlation correlation of 
the hard-sphere model at the activity z[l8[. It is worth 
mentioning that the second term of the rhs of Eq . (|A4j) 
corresponds to successive terms of a series expansion that 
can be resummed, and finally, one obtains that 



^ = ^M + , s lnfl + ^]-^) 



dv{2h*{r\z) - (h*(r\z)) 2 ) + 0(l/p 3 s ). (A8) 



V. A. Bakaev, Surface Science 564, 108 (2004). 

Z. Adamczyk, B. Siwek, P. Weronski, and E. Musial, Ap- [14 

plied Surface Science 196, 250 (2002). 

A. Dabrowski, Adv Colloid Int Sci 93, 135 (2001). [15 
Z. Adamczyk, K. Jaszczolt, A. Michna, B. Siwek, 
L. Szyk-Warszynska, and M. Zembala, Adv Colloid Int [16 
Sci 118, 25 (2005), ISSN 0001-8686. 

M. Jaroniec and R. Madey, Physical adsorption on het- [17 
erogeneous solids (Elsevier, 1988). 

W.Rudzinski and D. H. Everett, Adsorption of Gases on [18 
Heterogeneous Surfaces (Academic Press, 1992). 
R. D. Johnson, Z. G. Wang, and F. H. Arnold, J. Phys. [19 
Chem. 100, 5134 (1996). 

J. Talbot, G. Tarjus, and P. Viot, Phys. Rev. E 76, [20 
051106 (2007). [21 
G. Oshanin, O. Benichou, and A. Blumen, J. Stat. Phys. [22 
112, 541 (2003). 

G. Oshanin, O. Benichou, and A. Blumen, Europhys. [23 
Lett 62, 69 (2003). 

X. Jin, N. H. L. Wang, G. Tarjus, and J. Talbot, J. Phys. [24 
Chem. 97, 4256 (1993). [25 
C. Oleyar and J. Talbot, Physica A 376, 27 (2007). 
E. Ostuni, B. A. Gryzbowski, M. Mrksich, C. S. Roberts, 



and G. M. Whitesides, Langmuir 19, 1861 (2003). 

E. Liithgens and A. Janshoff, ChemPhysChem 6, 444 

(2005). 

M. Weight and A. K. Hartmann, Phys. Rev. Lett. 26, 
6118 (2000). 

M. Weight and A. K. Hartmann, Phys. Rev. E 63, 056127 
(2001). 

M. L. Rosinberg, G. Stell, and G. Tarjus, J. Chem. Phys. 
100, 5172 (1994). 

J.-P. Hansen and I. R. McDonald, Theory of Simple Liq- 
uids (Academic Press, 1976). 

G. Stell, Phase transitions and critical phenomena (Aca- 
demic Press, 1976), p. 205. 
L. Tonks, Phys. Rev. E 50, 955 (1936). 
S. Torquato, Phys. Rev. E 51, 3170 (1995). 
M. D. Rintoul, S. Torquato, and G. Tarjus, Phys. Rev. 
E 53, 450 (1996). 

P. Viot, P. R. V. Tassel, and J. Talbot, Phys. Rev. E 57, 
1661 (1998). 

X. Z. Wang, Phys. Rev. E 66, 031203 (2002). 

J. Quintanilla and S. Torquato, Phys. Rev. E 54, 5331 

(1996).